Closing the Loop on Bond Selective Chemistry Using Tailored Strong

Jun 18, 2002 - (OCE) with an emphasis on closed-loop control methods using near-infrared fs pulses. Because the experiments are in the strong field re...
0 downloads 0 Views 295KB Size
© Copyright 2002 by the American Chemical Society

VOLUME 106, NUMBER 27, JULY 11, 2002

FEATURE ARTICLE Closing the Loop on Bond Selective Chemistry Using Tailored Strong Field Laser Pulses R. J. Levis* Department of Chemistry, Wayne State UniVersity, Detroit, Michigan 48202

H. A. Rabitz* Department of Chemistry, Princeton UniVersity, Princeton, New Jersey 08544 ReceiVed: September 12, 2001; In Final Form: January 28, 2002

Strong field, closed-loop control of gas-phase photochemical reactivity is the focus of this article. The control of chemical reactivity is now possible using tailored laser pulses to circumvent previous laser bandwidth limitations. As an illustration of this capability, ketone rearrangements and dissociation reactions are considered. To introduce the experiments we discuss both optimal control theory (OCT) and optimal control experiments (OCE) with an emphasis on closed-loop control methods using near-infrared fs pulses. Because the experiments are in the strong field regime, we present the current state of the understanding of the electronic and nuclear photophysical processes that occur when polyatomic molecules are subjected to laser intensities ranging between 1013 and 1015 W cm-2. Photoelectron spectroscopy measurements are presented that begin to elucidate the control mechanisms. These delineate the order of the multiphoton process, the presence of transient shifting of excited electronic state energies (on the order of 5 eV), and the phenomena of lifetime broadening of electronic states. Recent experiments probing the energy partitioning to nuclear modes are presented with an emphasis on detecting the final kinetic energy of fragment ions. The advances in laser pulse shaping technology slaved to pattern recognition learning algorithms have opened up the prospect of studying the dynamics and chemical manipulation of virtually any system that can be introduced into the closed-loop apparatus. Rather than operating under the limitation of finding the molecule to suit the laser capabilities, the closed-loop learning control procedure operating in the strong field regime now makes it possible to merely tailor the control laser to meet the molecule’s dynamical capabilities in keeping with the chemical objectives. The prospects are very bright for exploring chemical reactivity with these tools.

I. Introduction The interaction of molecules with tailored laser pulses in the strong field regime is under active exploration for optical control of chemical reactivity. For the processes described in this article, the strong field regime is reached at laser intensities in excess of 1012 W cm-2, where substantial Stark shifting, polarization, and disturbance of the field free electronic states occurs to produce a quasi-continuum of new states in the molecule. A calibration for the magnitude of the influence of an intense laser

on a molecule can be obtained by calculating the maximum amplitude of the electric field vector of the laser beam using

Eo ) (I/oc)1/2

(1)

where I is the intensity of the radiation, o is the vacuum permittivity, and c is the speed of light. For example, the easily obtained intensity of 1014 W cm-2 corresponds to Eo ) 2.75 V/Å. The result of the interaction of such a high electric field

10.1021/jp0134906 CCC: $22.00 © 2002 American Chemical Society Published on Web 06/18/2002

6428 J. Phys. Chem. A, Vol. 106, No. 27, 2002 where an initial quantum state |ψi〉 is steered to a desired final state |ψf〉 via interaction with some external field. As a problem in quantum control, the goal is typically expressed in terms of seeking a tailored laser electric field (t) that couples into the Schro¨dinger equation

∂ ip |ψ〉 ) [H0 - µ ‚(t)]|ψ〉 ∂t

Figure 1. Schematic of a diatomic molecule interacting with an instantaneous, strong electric field. In this case the electric field strength is on the order of the field of the H2 molecule 1 Å from the nucleus. The process shown represents tunneling of the electron into the continuum.

with a molecule is schematically shown in Figure 1 where the one-dimensional electrostatic potential energy surface of a diatomic molecule is modified by a laser pulse of approximately 1 V/Å, a field strength that is on the order of the fields binding valance electrons to nuclei. The control of the strong fieldinduced near continuum using closed-loop methods has been used to influence gas-phase chemical reactions1 where the outcome of the interaction between the strong field laser and the molecule is employed to interactively discover the optimal time-dependent pulse.2 At first one might anticipate that the degree of chemical control using pulses of such intensity would be extremely limited due to the highly nonlinear processes induced in the molecule. However, because the pulse duration is short (∼50 fs), the excitation laser has the potential to limit the intuitively expected catastrophic decomposition to atomic fragments and ions. For example, in the strong field excitation of benzene,3 ionization of the parent species was exclusively observed up to intensities of 1014 W cm-2 with little induced dissociation. The observation of the single dominant channel (intact ionization) suggested that most of the possible final state channels (i.e., the large manifold of dissociative ionization states) may be suppressed in a welldefined, strong field intensity regime. Furthermore, at these intensities there is opportunity to substantially manipulate the molecular wave function (see Figure 1) with suitable shaping of the laser pulses to induce and manage photochemical reactivity and products. The advent of short pulse duration, intense lasers has led to the observation of many interesting strong field phenomena in atoms, molecules, and clusters including X-ray generation from high harmonics;4 above threshold ionization;5 above threshold dissociation;6 multiple electron emission from molecules;7 intact ionization of large polyatomic molecules;3,8,9 forced molecular rotation in an optical centrifuge;10 production of extremely high charge states from molecular clusters;11 production of highly energetic ions;12 and neutrons from clusters.13 A clear picture of the excitation mechanisms in the strong field regime is now emerging.14 The use of this picture to understand the recent strong field, closed-loop control experiments is one focus of this review. While the use of strong fields to control chemistry is quite new, the area of coherent control research has broad foundations.15-17 The essence of the control concept in terms of optical fields and molecules is captured by the following transformation goal:

|ψi〉f|ψf〉

(2)

(3)

through the dipole µ. This Born-Oppenheimer picture can be expanded to explicitly consider the electrons and nuclei (cf., the discussion in section II). Regardless of the necessary level of Hamiltonian detail, the general mechanism for achieving quantum control is through the manipulation of constructive and destructive quantum wave interferences. The goal is to create maximum constructive interference in the state |ψf〉 according to eq 2, while simultaneously achieving maximum destructive interference in all other states |ψf′〉,f′ * f at the desired target time T. A simple analogy to this process is the traditional double slit experiment.18 However, a wave interference experiment with two slits will lead to only minimal resolution. Thus, in the context of quantum control, two pathways can produce limited selectivity when there are many accessible final states for discrimination. Rather, a multitude of effective slits should be created at the molecular scale in order to realize high quality control into a single state,19 while eliminating the flux into all other states. The requirement of optimizing quantum interferences to maximize a desired product leads to the need for introducing an adjustable control field (t) having sufficiently rich structure to simultaneously manipulate the phases and amplitudes of all of the pathways connecting the initial and final states. As will be seen, construction of such a pulse is currently possible in the laboratory using the technique of spatial light modulation.20,21 However, calculation of the time-dependent electric fields to produce the desired reaction remains a problematic issue for chemically relevant reactions. Unfortunately, the Hamiltonian at the Born-Oppenheimer level remains largely unknown for polyatomic molecules, and this severely limits the ability to perform a priori calculations at the present time. Even if the field free molecular Hamiltonian were known, the highly nonlinear nature of the strong field excitation process effectively removes all possibility of calculating an appropriate pulse shape in this regime. Thus we are left with the following conundrum: If the design can be carried out reliably, then the physical system will likely not be of much interest, while for interesting physical systems, reliable designs cannot be performed. The method of closed-loop control for laser-induced processes2 offers a way to surmount our lack of knowledge of the Hamiltonian to find appropriate pulse shapes, (t). In closedloop operations, the molecule, the laser pulse shaper, and a pattern recognizing learning algorithm form the elements for repeated cyclic operation to teach the laser how to control the molecules. A schematic of this process is shown in Figure 2. This procedure, especially in the strong field regime, provides the only general means at the present time to deduce laser pulse shapes that can successfully manipulate molecular dynamics phenomena. The method is general because any molecule can be excited in the strong field regime using a nominally 800 nm pulse, and closed-loop methods provide a means to determine the optimal time-dependent, strong field excitation to produce a specific target state. The quantum system, upon each cycle of the loop, is replaced by a new one, thereby (a) avoiding the need for ultrafast computations, electronics, and laser switching, and (b) eliminating any concerns about the observation process

Feature Article

J. Phys. Chem. A, Vol. 106, No. 27, 2002 6429

Figure 3. Potential outcomes of the interaction of intense laser radiation with a molecule. At the present time the wavelengths used for the interaction range between 10µm and 200 nm. The wavelengths employed in the studies reported here range between 750 and 850 nm with intensities of 1013 to 1015 W cm-2. Figure 2. Representation of the concept of closed-loop control. In this case there are four possible outcomes shown for the interaction of the laser pulse with the molecule. The desired distribution of products is first input into the algorithm at the upper left of the Figure. The program creates an initial laser pulse shape that interacts with the sample and yields a product distribution. Based on experimental measurement of the distribution (typically in combination with several other experiments) the algorithm creates a new pulse that yields a new distribution. The system loops iteratively until the desired level of control is exerted.

including both electronic and nuclear dynamics. Principles of quantum optical control, and especially closed-loop learning control with tailored femtosecond laser pulses, are discussed49 in section III. Laboratory laser control of atomic and molecular processes in the strong-field regime is the subject of section IV. Finally, section V considers future trends and possible new applications of closed-loop laser control of molecular dynamics phenomena. II. Molecules in Intense Laser Fields

disturbing the actual dynamics. In the experiments described in this article, the phase and amplitudes of the component frequencies of a 40 fs pulse are the control variables, and the resultant mass spectrum is the observable employed to evaluate the fitness of the pulse shape. Optical or other means of detection could also be employed for a feedback signal. The closed-loop learning procedure for teaching lasers to control quantum systems has now been demonstrated in many diverse investigations. Those performed to date employ closedloop control of a laser pulse shaper to optimize a desired process. These experiments include adaptive pulse compression22,23 and control of pulse phase,24 manipulation of pure rotational25 and vibrational26 dynamics in diatomics, one and two photon transitions in atoms,27,28 creation of specific wave functions in Rydberg atoms,29 and generation of high harmonics in Ar gas.30 In the case of controlling chemical processes, a number of experiments have been performed. These include schemes employing low lying resonances to maximize fluorescence from a dye molecule in solution,31 generation of specific photochemical fragments from organometallics32,33 and alkali clusters,34 maximization of stimulated Raman signal from methanol in solution,35 and the optimization of coherent antistokes Raman emission.36 Most recently, shaped, strong field laser pulses have been employed to enable the control of photodissociation processes in organic molecules.1,37 The combination of closedloop operations with strong field laser control has opened the door to the ready control of chemically interesting processes. This paper seeks to define the emerging area of strong field control of chemical reactivity using closed-loop, tailored light pulses. To do this we review several relatively new areas of research including closed-loop optimization and strong field processes. We also review the experimental linkage of these two areas through spatial light modulation of intense laser radiation. Experiments using intense near-infrared laser pulses will be considered here. The paper is organized into the following sections. Section II reviews the processes resulting from the interaction of molecules with strong laser fields;

When a molecule interacts with an intense laser pulse, a number of product channels may be accessed. Some of the potential outcomes are listed in Figure 3 where coupling into the nuclear, electronic, and nonlinear optical channels are delineated. Initial intuition suggested, incorrectly, that intense, short duration laser pulses interacting with polyatomic molecules would result primarily in multiphoton dissociation as shown in the first channel. Early experiments using intense nanosecond, picosecond, and femtosecond pulses provided ample evidence for the second and third coupling channels in Figure 3, which may be described as dissociative ionization and Coulomb explosion,38 respectively. Pulses of femtosecond duration have been shown to couple into electronic channels, resulting in ionization without nuclear fragmentation for molecules such as benzene and naphthalene.3 In such experiments the energy in excess of the ionization potential (up to 50 eV!)39,40 couples mainly into the kinetic energy of the photoelectron. In terms of control experiments, the ability to produce intact ions at such elevated laser intensities suggested the possibility that intense lasers could be used to guide the dynamics of a molecule into a channel other than catastrophic decomposition. Molecules interacting with intense laser fields may also convert the fundamental of the excitation laser into higher harmonics.41 This review will focus on the channels of dissociative ionization and molecular ionization listed in Figure 3. The relative importance of each product channel shown in Figure 3 is dictated by the Hamiltonian for the moleculeradiation system. Our understanding of the Hamiltonian for polyatomic molecules in general, and the more complex Hamiltonian for the interaction between strong fields and molecules in particular, is rather limited at the present time.9 One would like to have high quality time-dependent calculations to model the strong field interaction, but these are simply intractable with current computational technology. Calculations for simple systems containing up to three protons and one or two electrons have been performed, and these systems are reasonably well understood.42-44 For polyatomic systems, the

6430 J. Phys. Chem. A, Vol. 106, No. 27, 2002

Figure 5. Schematics of the photoelectron spectrometer and the timeof-flight ion detector used for measuring the kinetic energy distribution and molecular weight of the product ions.

Figure 4. Schematic of the structure-based model for representing molecules in intense fields. The presentation in the left-hand panel is the zero-range model where only the ionization potential of the system is employed in calculations. The presentation in the right-hand panel represents the use of the electrostatic potential of the molecule in determining an appropriate one-dimensional rectangular well to represent the spatial extent of the system. To compare the models, an electric field of 1 V/Å is superimposed on each potential to reveal the barrier for tunnel ionization.

number of degrees of freedom is too large for first-principles calculations. Thus, simple models have been employed to gain some insight into the mechanisms of interaction between intense laser pulses and atoms. There is a hierarchy of models for representing molecules interacting with intense laser fields. The earliest viewed the potential energy of interaction between the electron and the core as a delta function having a single state at the ionization potential of the system (called a zero-range potential).45 Subsequently, a Coulomb potential was employed for calculations in atoms.46,47 This was followed by a rectangular potential for molecules defined within the context of the structure-based model as shown in Figure 4.9,39,48-50 The rectangular potential approximates the delocalization of electrons over the length scale of the molecular dimension by defining the width of the well to be equal to the characteristic length of the molecule. The characteristic length is defined as the largest distance between classical turning points in the three-dimensional electrostatic potential energy surface at the ionization potential of the molecule. The height of the rectangular well is the ionization potential of the molecule. A further advance incorporated time dependence into the radiation-molecule interaction to go beyond the quasi-static regime.51

The experiments reviewed in this article focus on both photoelectron and photoion measurements to determine the basic phenomenology of strong field excitation of polyatomic species and to test theoretical models. The apparati used to measure the photoelectron and photoion distributions are shown in Figure 5. The electron kinetic energy distributions have been measured as a function of molecular structure and laser intensity. The photoelectron distribution provides a snapshot of the intense laser-molecule interaction during excitation because the time scale for photoelectron ejection is short (∼fs) in comparison with the time scale for photoion decomposition (∼ps). The latter provides information about the final state distribution of the laser-molecule interaction. The photoion experiments reviewed here include measurements of the ion mass distribution and ion kinetic energy distribution. The photoion kinetic energy distribution measurements compliment the photoelectron measurements regarding the final state energy partitioning after strong field excitation. A. Electronic Dynamics of Molecules in Intense Laser Fields. To describe the mechanisms of strong field control of chemical processes it is important to consider the influence of the intense laser field on electrons in the molecule. For instance, we will see that bound electrons can gain significant ponderomotive energy (∼1-5 eV) during the pulse, and eigenstates can shift by similar energies.52 In the case of the interaction of a laser pulse with a molecule, the appropriate starting point is the Hamiltonian for a multielectron system interacting with an electromagnetic field:

H)

Pc2 2M

+

1

z

1

z

∑ PI2 + m i>j)1 ∑ Pi ‚ Pj + V(x1..., xi, ...xc) + 2µ I)1 n

e µc

A(xc, t)

i)1 Z

Pi +

Ze2 2µc

2

A2(xc, t) (4)

where P is momentum,V is the potential energy as a function of position, Z is the nuclear charge, and A(xc,t) is the vector potential of the laser radiation. The first four terms describe the field free motion of the system. The last two terms describe the effect of the laser radiation on the population of eigenstates and corresponding shifts in the eigenstates of the system. In the electric field gauge the last term becomes

Ze2E2(xc, t)/(2 µω2)

(5)

where E is the electric field of the laser, and ω is the frequency of the laser. The average of this term over the period of

Feature Article

J. Phys. Chem. A, Vol. 106, No. 27, 2002 6431

Figure 6. The effect of various terms in the Hamiltonian for a charged particle in an oscillating electromagnetic field is shown. The ionization potential of the system remains unchanged by the A2 term as all states are raised equally. The A‚P term lowers the ground state of the system by an amount equal to the A2 term plus an additional amount due to the induced polarization of the system. The net result is an increase in the ionization potential by an amount approximately equal to the ponderomotive potential of the laser pulse.

oscillation for linearly polarized light is

Up ) (Ze2E02)/4 µω2)

(6)

where E0 is the amplitude of the electric field. Up is known as the ponderomotive potential. In strong fields this term shifts all eigenstates upward in energy equally by Up. A differential shifting of eigenstates results from the A‚P term. To first and higher order, the A‚P term may used to describe allowed transitions of amplitude between eigenstates. To second and higher order, this term will describe differential shifting of the eigenstates. The magnitude and sign of the shift of a given state are dependent on the wavelength and the electronic structure of the system. Pan et al.53 have derived expressions for the shifting of the ground state and Rydberg/continuum states of a model system. A lowest nonvanishing order perturbation theory treatment53 yields the ground (∆Eg) and Rydberg level (∆ER) energy shifts as

∆Eg ) -

Ze2E02 4µω

2

- 1/2 RE02

∆ER ≈ 0

(7) (8)

where R is the ground-state polarizability. The first term in eq 7 is the negative of the ponderomotive potential Up. The second term is equivalent to the dc Stark shift. This treatment is valid when the ground state is deeply bound and separated from adjacent eigenstates by many times the photon energy, hν (the low-frequency approximation). This is valid for most atoms and molecules investigated with near-infrared or longer wavelength light. High lying bound states and all continuum states experience no A‚P shift, whereas deeply bound states of the atom experience a much greater, negative shift.54 The pertinent shift in the states as a function of the terms in the Hamiltonian in the long wavelength limit is summarized in Figure 6. The laser intensities employed in recent high field experimental manipulation of chemical reactivity range up to 5 × 1014 W cm-2. This corresponds to ponderomotive shifts up to 10 eV with similar shifts in the separation of the ground and excited-state potential energy levels. The laser employed in these investigations has a period of 2.5 fs and an envelope with fwhm of 60-170 fs corresponding to at least a several hundred significant oscillations in the electric field vector interacting with

the molecule. The states of the molecule undergo an associated oscillation in the splitting between energy levels that may result in periodic excitation on a time scale of the period of the laser. This dynamic shifting of energy levels implies that there will be transient field-induced resonances (or Freeman resonances).55 Evidence for these resonances in the case of molecules has been obtained by measuring the strong field photoelectron spectroscopy of a number of molecules including acetone, acetylene,52 water, benzene, and naphthalene.56 The oscillatory nature of the intense laser excitation also leads to above threshold ionization (ATI) peaks in the photoelectron spectrum.5 These are denoted by peaks spaced by the photon energy extending to many photons above the minimum number required for ionization. In the case of acetylene,52 strong field photoelectron evidence has been found for substantial shifting (1-4 eV) of the 4p series of Rydberg states to attain resonance. Peaks observed in strong field photoelectron spectra can be assigned via transient shifting of states by an amount up to the ponderomotive potential of the laser. The method developed to measure and assign the spectra is called field-induced resonance enhanced multiphoton ionization (FIRE MPI). To assign the strong field spectra one first calculates Up′(I), the variable ponderomotive shift required to bring a given candidate state into l photon resonance52 (see Figure 6 for the field dependent shifting of states). This virtual ponderomotive shift is

Up′(I) ) lhν - Estate

(9)

where Estate is the energy of the given state under field-free conditions. Up′(I) is allowed to shift up to the maximum ponderomotive potential of the laser, Up, because resonance may occur at any intensity within that range. The IP of the system at the instant of resonance is shifted simultaneously to some higher (intensity dependent) value given by

IP′(I) ) IP + Up′(I)

(10)

The kinetic energy of the photoelectron generated by absorbing m additional photons above the l photon resonance is then

Efeature ) (l + m)hν - IP′(I)

(11)

In the limit where Up′ ) 0, eq 5 reduces to the conventional REMPI condition, Efeature ) (l + m)hν - IP, where lhν ) Estate. However, in strong laser fields, Estate may shift into l photon resonance, giving rise to the transient features observed in FIREMPI. All potential l-photon resonances are analyzed to determine those generating photoelectrons of appropriate kinetic energy. Such field-induced shifting of intermediate states provides a powerful mechanism for strong field coupling to molecules and may be responsible in part for the strong field control mechanism. In the case of acetylene, the measurements also demonstrated that strong field photoelectron spectroscopy could detect states that had been theoretically predicted but were not detected using nanosecond resonance-enhanced multiphoton ionization because of short-lived states. Transient resonances are not the only strong field processes induced upon molecular eigenstates during intense laser. The electric field can also broaden molecular states through a lifetime mechanism. Lifetime broadening is expected for any mechanism that causes decay of population from a given state, including, for example, ionization, nonadiabatic effects, and dissociation. In the static limit, an electric field superimposed on any system can result in tunneling. The tunneling rate may be calculated using the WKB approximation (modeling the system as one-

6432 J. Phys. Chem. A, Vol. 106, No. 27, 2002

Figure 7. A schematic of the field-induced broadening resulting from a decreased lifetime of ground and excited states from ionization processes. Also shown is the field-induced shifting of the ground state to lower energy as a result of the intense laser pulse. Both of these processes contribute to an increase in the effective bandwidth in the excitation process.

dimensional) where the rate is given by

∫rr

w ) exp{- 2

2

[2(IP - V(r))]1/2 dr}

(12)

1

where the limits of integration are defined by the path length for tunneling, as shown in Figure 4. The lifetime of the molecule in the neutral state is given by the inverse of the tunnel ionization rate. Thus, an upper limit is placed on the lifetime of the neutral state which may then be related to the uncertainty of the state energy by the Heisenberg relation

∆E∆ t ) h or ∆Ew-1 ) h

(13)

Lifetime broadening enhances the opportunity for excitation and may be thought of as a mechanism for increasing excitation bandwidth. Here, the excitation bandwidth is the combination of the laser bandwidth and the width of the field-induced quasi continuum. In the weak field regime, the excitation bandwidth is given almost exclusively by the bandwidth of the exciting laser. A long duration radiation source, such as a nanosecond laser, will have a bandwidth of ∼µeV, while a 100 fs duration laser will have a bandwidth of ∼50 meV. In the weak field case the excitation scheme is necessarily limited to states that fall within the spectral range of the excitation source and possibly low harmonics of that source. In the strong field case there is an opportunity to increase the bandwidth of the excitation laser by widening the eigenstate to a band of perhaps several eV by the lifetime broadening mechanism. A schematic of the combined effects of state shifting and lifetime broadening is shown in Figure 7. These strong field effects provide an attractive regime to consider for molecular control, as one no longer needs to search for a molecule that suits the finite laser frequencies available in the laboratory. Rather, the laser pulse may be tailored to suit virtually any molecule. Evidence for the broadening of eigenstates during the strong field excitation process can be found in the photoelectron measurements for acetylene,50 benzene, and naphthalene.56 In each of these molecules the photoelectron spectra contain a welldefined series of features that can be assigned using the method of FIRE MPI as described previously. As the laser intensity increases above that required for detection of photoelectrons,

Figure 8. Strong field photoelectron spectrum for benzene shown on an energy axis that includes the photons necessary to induce ionization. The photoelectron spectrum was obtained using 2 × 1014 W cm-2, 800 nm radiation of duration 80 fs. The quantum energy of the photons are shown to scale and indicate that 10-20 photons are available to drive excitation processes in the strong field excitation regime. In addition, uncertainty broadening of the pulse will also produce a distribution of allowed photon energies that approaches the photon energy when multiphoton processes of order 10 are approached.

the features begin to broaden. In general at an intensity of roughly 1 order of magnitude larger than the ionization threshold, the discrete features are smeared into a continuum. This implies that for these highly nonlinear processes broadening on the order of several eV occurs rapidly above the threshold for ionization, perhaps through the lifetime mechanism. The fact that intact ions are observed in the mass spectra at these elevated intensities suggests that ionization of dissociated products is not responsible for loss of the features. A similar broadening of eigenstates has been observed at constant laser intensity in the series benzene, naphthalene, and anthracene at constant laser intensity where the characteristic length of the molecule increases.39 In the case of benzene, having the smallest characteristic length and hence the largest barrier to tunnel through, there are several series of observable features. In the case of naphthalene, having a larger characteristic length, there are discrete features superimposed on a feature having a broad distribution of energies. Anthracene reveals no evidence for well-resolved peaks within the broad photoelectron distribution. These observations suggest that the lifetime broadening scales with increasing characteristic length. An important consideration for the control of chemical reactivity in the strong field regime is the order of the multiphoton process during excitation. This order indicates the maximum number of photons that are available to drive a chemical reaction. Some indication of the number of photons involved in the strong field excitation process can be gleaned from measurements of strong field photoelectron spectra. Figure 8 displays the photoelectron kinetic energy distribution for benzene with the energy axis rotated by 90 degrees. The energy scale has been offset to include the energy of the ground and ionization potential of the molecule in the absence of the strong electric field. The arrows on the figure represent the photons involved in both exceeding the ionization potential and in

Feature Article creating the above threshold ionization photoelectron distribution. At least six photons are required to surmount the ionization potential of benzene. Recall that in the presence of the strong electric field, the ionization potential will increase by an amount greater than the ponderomotive potential, further increasing the actual number of photons involved in the excitation process. At the intensity of 1014 W cm-2 in this measurement, on the order of 10 photons may be absorbed to induce the photoelectron spectra shown. Including the photons required to reach the ionization potential, this means that approximately 20 photons may be involved in the excitation process. With the shaped pulses used in the experiments described in section IV, the intensities are lower and on the order of 10 or fewer photons are likely involved in the excitation process. Several other methods have been developed to predict the ionization probability of molecules. One is based on discretizing a molecule into a collection of atomic cores that individually interact with the strong laser field and emit electrons.57 In this model, a carbon atom, for instance, is represented by an atom with an effective potential. The ionization probability is then a function of the individual ionization probabilities from atoms with opportunity for quantum interference during the ionization event. Unfortunately, the method must be parameterized for each molecule at the present time. The second method under development employs S-matrix57 theory to calculate the ionization probability for atoms and now molecules. This method focuses on the interference of the outgoing electron wave. Predictions about relative ionization probabilities are based on the symmetries of the highest occupied molecular orbital. B. Nuclear Dynamics of Molecules in Intense Laser Fields. The response of a molecule to a time-dependent electric field is the means by which chemical reactivity is controlled in these experiments. In the case of weak laser fields, the response can be calculated with reasonable accuracy.58-60 In the case of strong fields, the situation is much more complex but the dynamical possibilities are much richer. In principle, the nuclear dynamics in strong laser fields could be determined using exact numerical solutions of the time-dependent Schro¨dinger equation. Such solutions are possible only for the simplest of molecules at the present time.42-44 In fact, the bulk of such simulations have been performed using a one-dimensional model for the H2+ system.61-63 These calculations show the presence of non-BornOppenheimer electron-nuclear dynamics. Since the nuclei move considerably on the time scale of the laser pulse, electronic modes are necessarily coupled with nuclear modes. Three distinct final states have been observed in strong field (no pulse shaping) mass spectra of polyatomic molecules: production of intact molecular ion, ionization with molecular dissociation, and removal of multiple electrons to produce Coulomb explosion.9 The hallmark of the latter process is production of ions substantial (>5 eV) kinetic energy. The presence of Coulomb explosion has been shown to depend on charge resonanceenhanced ionization64 (CREI) which becomes the dominant mechanism at large critical internuclear distances. Interestingly, the production of high charge states in molecular clusters can be controlled using pump-probe excitation schemes.65 At intensities that are lower than the threshold for multielectron ionization, the majority of molecules display some fraction of intact ionization. This phenomenon is not expected intuitively because the ionization processes are not resonant with low order multiples of the fundamental frequency, implying that intense pulses must be employed for excitation. Nonetheless, many molecules have been investigated to date and all appear to provide some degree of intact molecular ionization when 800

J. Phys. Chem. A, Vol. 106, No. 27, 2002 6433

Figure 9. Retarding field measurement of the H+ ion kinetic energy distributions arising from benzene, naphthalene, anthracene, and tetracene after excitation using 2 × 1014 W cm-2, 800 nm radiation of duration 80 fs. The measurements reveal that as the characteristic length of the molecule increases, the cutoff energy increases monotonically.

Figure 10. Time-of-arrival distributions for the H+ ions for benzene, naphthalene, anthracene, and tetracene after excitation using 2 × 1014 W cm-2, 800 nm radiation of duration 80 fs. The time of arrival distributions were measured by allowing the ions to drift in a field free zone of length 1 cm prior to extraction into the drift tube. In this experiment, earlier arrival times denote higher kinetic energies.

nm excitation is employed. The mechanism behind this ionization appears to involve suppression of ladder switching coupled with coherent excitation of electronic modes. The state of this subject has been reviewed recently.9,66,67 To measure the amount of energy that may couple into the nuclear degrees of freedom during the intense laser excitation event, we have investigated40 the kinetic energy release in H+ ions using both time-of-flight and retarding field measurements. A typical time-of-flight mass spectroscopy apparatus employed to make such measurements is shown in Figure 5. In the series benzene, naphthalene, anthracene, and tetracene, the most probable kinetic energy in the measured distributions was observed to increase as the characteristic length of the molecules increased as shown in Figure 9. The corresponding retarding field measurements are shown in Figure 10. Again the coupling into nuclear degrees of freedom was observed to increase in the larger molecules. The most probable kinetic energies increased from 30 V for benzene to 60 V for tetracene when a 1.2 × 1014 W cm-2 laser excited the molecules. In terms of

6434 J. Phys. Chem. A, Vol. 106, No. 27, 2002 providing an enabling capability for strong field control, these results suggest that up to 80 photons may be involved in the excitation process when a molecule such as tetracene is excited under strong field conditions. A general observation after ionization of large polyatomic molecules is the measurement of an enhanced degree of dissociation as the length of the molecule increases. This was first attributed to field-induced effects3 without a quantitative model. Recently, a strong field nonadiabatic coupling model has been introduced to account for the enhanced coupling into nuclear modes in molecules with increasing characteristic length.51 This excitation is akin to plasmon excitation where the precise energy of the resonance depends on the coherence length and binding energy of the electrons and the strength and frequency of the driving field. The model considers the amplitude of electron oscillation in comparison with the length of the molecule. If the amplitude of oscillation is small, the molecule may first absorb energy nonresonantly and then ionize from the excited states. The amplitude of the electron oscillation in an laser field is given by aosc ) E/ωL2. In the event that the aosc< L, where L is the characteristic length of the molecule, the electron gains ponderomotive energy from the laser. Given an energy level spacing of ∆ο, the probability of nonadiabatic excitation within the Landau-Zener model becomes exp(π∆ο2/4ωLEL). As described,45 the threshold for nonadiabatic excitiation (when ∆ο2 ) ωLEL) of a 4 eV transition for a system having L ) 13.5 Å with 700 nm radiation occurs at 5.6 × 1012 W cm-2. This theory implies that the probability for exciting nuclear modes in large molecules with delocalized electronic orbitals increases monotonically with characteristic length as observed experimentally.3,51 The theory also suggests that intact molecular ionization will increase with increasing excitation wavelength for large molecules, and this has been confirmed.51 Whether the nonadiabatic excitation can be controlled remains an open question at the present time. The present successes1,32,37 in controlling chemical reactivity suggest that nonadiabatic processes either are not significant or that the closed-loop control method is able to effectively deal with this excitation pathway. III. Theoretical Concepts for Controlling Molecular Dynamics Phenomena A. General Considerations. The material in section II spelled out the phenomena and mechanisms operative when strong laser fields interact with polyatomic molecules. The present section will introduce the formal concepts and principles underying the control of molecular dynamics using such laser-induced processes. Attempts at controlling molecular-scale phenomena with lasers have a long history,68 going back to the early 1960s. It is useful to freshly examine the basic objectives and desires while considering the special features provided by operating in the strong field regime. Perhaps the most important aspect of operating with strong fields is the ability to move the molecular energy level resonances about, as necessary, to cooperate with the laser capabilities and thereby create molecular electronicnuclear wave packets with great flexibility.1 An essential feature of this process is the effective broad bandwidth provided the strong field interactions with the molecule. Analogous broad bandwidth control capabilities may also emerge from other laser technologies (e.g., locking together multiple lasers operating at distinct frequencies) in the future.69 Regardless of the control field characteristics, a basic goal of all chemical experiments is to achieve the best possible outcome (e.g., selective manipulation of reactivity). Thus an optimization process is a desirable way to manipulate molecular-

scale phenomena, thereby laying the foundation for introducing optimal control theory70-72 (OCT) as well as the allied realization of optimal control experiments73 (OCE). The notions of OCE, and especially its practical closed-loop implementations,2,74-84 have roots in OCT, and both procedures share some common algorithmic features. Given the general goal of steering the dynamics of the molecular system, the next consideration is how to identify the appropriate laser fields to meet the posed objectives. Some 40 years ago, at the inception of laser control over reactivity, simple chemical intuition was thought to be sufficient for this purpose;68 the lack of significant positive results over the subsequent approximately 30 years speaks to the inadequacy of using intuition alone. Physical intuition will always play a central role; however, it needs to be channeled into the appropriate mathematical and laboratory frameworks to be useful. A traditional approach to discover appropriate laser fields for molecular control would be through theoretical design, followed by implementation of the design in the laboratory upon the actual molecular sample.70,72,73,85,86 This logic, folded in with the desire to achieve the best possible results, is the essence of OCT for attaining laser field designs. Although many practical difficulties may be encountered in executing such designs for interesting chemical systems (i.e., polyatomic molecules), OCT laid the foundation for OCE2,87 leading to the recent successful laser experiments on manipulating chemical reactivity in complex molecules1,32 and other systems.30,35,88,89 In addition, the largely informational inadequacies (i.e., lack of quantitative knowledge of the Hamiltonian) and computational difficulties plaguing OCT are not inherent. Algorithmic and other advances will surely lead to better design capabilities in the coming years. For all of these reasons, section IIIB will summarize the general concepts behind OCT. Although the capability of designing laser fields to achieve particular physical objectives is improving, a most interesting set of recent experiments,1,30-32,35,88,89 and especially those involving strong field manipulation of polyatomic molecules, have operated by performing OCE directly in the laboratory. The success of this detour around OCT fundamentally rests on the ability to perform high throughput laser control experiments,19 slaved to fast learning algorithms capable of operating at the apparatus duty cycle. In this fashion, patterns are rapidly identified in the control field f molecular response relationship emerging from each cycle of the closed-loop operations, thereby homing in on control fields that optimally achieve the desired physical objective. Notwithstanding the anticipated improvements in OCT and even the present ability to reliably perform laser field designs for certain “simple” chemical applications,73 it is reasonable to categorically state that, in the foreseeable future, closed-loop OCE will form the only practical means of achieving successful control of complex polyatomic molecules, especially with multiple product channels. Thus, section IIIC will express the general principles and procedures for closedloop OCE. B. Optimal Control Theory. A fundamental question to ponder before considering any control field design algorithms is whether any field exists that may lead to successful control in a particular quantum system. This question is addressed by a controllability analysis. Controllability tools are available to assess whether it is, in principle, possible to arbitrarily steer about the wave function90 and the more general time evolution operator91 in any given quantum system expressed in a finite dimensional basis. Although an affirmative answer to controllability of the wave function would immediately imply the

Feature Article ability to control any physical observable for the system, the control of a particular observable should be a less demanding task to assess and possibly achieve. Even a respectable level of partial controllability may be quite adequate for many applications. However, the tools have yet to be developed for assessing controllability of arbitrary physical observables. Putting aside fundamental issues of controllability, OCT forms a reliable design procedure to identify the best control field possible under a given set of conditions.70,72,73,85,86 The most comprehensive means for controlling a molecule undergoing complex dynamical evolution is through coordination of the controlling electromagnetic field with the molecule’s characteristics. The spectral content and temporal structure of the control field should be continuously alterable throughout the process. This tight coordination ensures that all of the dynamical capabilities (i.e., both electronic and nuclear) of the molecule can be exploited to best meet the chemical objectives. Given specified initial and final states of the molecule, as in eq 2, and any imposed restrictions on the field or molecular dynamics, the time-dependent control field required to meet the objective may be designed using OCT. This general formulation encompasses both the weak and strong field limits, and, in principle, is capable of discovering control methods based on two-pathway interference induced by monochromatic laser fields,92,93 the “pump-dump” techniques based on two ultrashort laser pulses,94,95 and control via stimulated Raman adiabatic passage.96,97 Optimal control theory has an extensive history in traditional engineering applications,98 but the quantum nature of molecularscale phenomena imposes special features. Consider a quantum system (e.g., a molecule), whose free evolution is governed by the Hamiltonian H0. The full Hamiltonian of the laser-driven system is H ) H0 - µ(t), with the dynamics prescribed by eq 3. A more complete picture with all electrons and nuclei specifically treated could be considered based on the Hamiltonian in eq 4. The goal of OCT is to design an electric field that will allow manipulation of the system dynamics in a desired way, subject to eq 3 being satisfied. A typical quantum control objective is to maximize the magnitude of the expectation value 〈y(T)|O|ψ(T)〉 of a specified observable operator O at the final time T. For example, O might be the flux operator associated with a reactive channel, with the control objective being maximization of the product yield in that channel. In practice, there may be multiple objectives involving distinct observable operators corresponding to the desire to simultaneously manipulate several physical aspects of the same system (e.g., control the fate of multiple bonds in a polyatomic molecule). In addition, there may be costs or constraints on the form, magnitude, frequency, or other characteristics of the control field. These various objectives and constraints will often be in competition with each other. This recognition motivates posing the control design problem as an optimization attempting to strike a balance between the competing physical goals. Balancing such competition is an essential feature of OCT. The physical objectives are expressed collectively in a cost functional J[ψ(t),(t)], dependent on the evolving wave function70 (or density matrix,99 if appropriate), the target states or expectation values, any constraints, and the electric field. Physical input, often guided by intuition, will enter through the form and relative weight given to the different terms in the cost functional. The cost functional J is optimized with respect to the control field (t), to yield the best possible control performance in balance with any other competing factors. Consider, for example, the case of maximizing the expectation value 〈y(T)|O|ψ(T)〉 of a positive definite operator O at the target

J. Phys. Chem. A, Vol. 106, No. 27, 2002 6435 time T, while minimizing the laser field fluence. In this circumstance, the cost functional may take the form

∫0T [(t)]2 dt T 2J[∫0 〈χ(t)|ip∂t - H|ψ(t)〉 dt]

J ) 〈ψ(T)|O|ψ(T)〉 - R0

(14)

Here, ∂t ) ∂/∂t,R0 is a positive parameter chosen to weight the significance of the laser fluence, |ψ(t)〉 is the system wave function, and |χ(t)〉 is a Lagrange multiplier introduced to ensure satisfaction of the Schro¨dinger equation in the design process. Requiring that the first variation of J with respect to |ψ(t)〉, |χ(t)〉, and (t) satisfy δJ ) 0 will give equations for the wave function, Lagrange multiplier, and optimized laser field:70,72

ip∂t|ψ(t)〉 ) H|ψ(t)〉,|ψ(0)〉 ) |ψ0〉

(15)

ip∂t|χ(t)〉 ) H|χ(t)〉,|χ(T)〉 ) O|ψ(T)〉

(16)

(t) ) - J〈χ(t)|µ|ψ(t)〉/R0

(17)

Here, |ψ0〉 is the initial state of the quantum system. Numerical solution of the above equations will give the desired optimal control field, although this often is a problem of significant computational complexity. Specifically, the accurate solution of the many-dimensional Schro¨dinger equation in eqs 15 and 16 poses a significant challenge even for cases with a few atoms. Equations 15-17 will generally have multiple solutions corresponding to a family of locally optimal control field designs.100 Various iterative algorithms have been developed for the calculation of optimal control fields and many numerical examples have demonstrated quantum optimal control of molecular-scale phenomena, (e.g., rotational, 101 vibrational,102 electronic,103 reactive,104,105 and other processes.106 In addition to achieving a balance among the physical objectives, the OCT design process may also include the goal of achieving the objectives while simultaneously having the process be as robust as possible to laser field errors or Hamiltonian uncertainties.107 The many OCT simulations performed in recent years have produced physical insight into the control of quantum phenomena. However, all of these efforts have been carried out with relatively simple systems or simple models of complex systems. Notwithstanding this comment, the most important result coming from the various OCT simulations is that successful control fields exist, capable of providing high quality molecular manipulation to meet many physical objectives. The significance of this conclusion stands, regardless of the fact that it is drawn from models of molecules and other systems. The ability to perform OCT will surely improve in the coming years, and it should continue to provide at least physical insight into virtually all control applications. C. Optimal Control Experiments. Optimal control theory has proved to be a valuable theoretical tool for exploring coherent laser manipulation of quantum systems. However, attaining laboratory-significant OCT field designs requires precise knowledge of the system Hamiltonian, which often is not available for the most interesting molecules or materials, even for those of modest complexity. The quality of laser field designs is also limited by the ability to accurately solve the design equations. To appreciate the significance of these comments, recall that quantum control relies on the often delicate manipulation of matter wave interferences through the proper tuning of laser phases and amplitudes. It is reasonable to expect that there will be only limited tolerance to inevitable field design errors.

6436 J. Phys. Chem. A, Vol. 106, No. 27, 2002 A crucial step toward attaining high-quality laboratory laser control of molecular-scale phenomena was the introduction of adaptive control techniques, initially suggested by Rabitz and Shi in 1991,87 and elaborated on by Judson and Rabitz in 19922 and in subsequent investigation.74-76,78-81 In this OCE approach, known as learning control, a loop is closed in the laboratory around the quantum system, with results of the control field f observable outcomes used to evaluate the success of the candidate applied laser field designs and to refine them, until the control objective is reached as best as possible with an optimal field. In learning control, a new molecular sample is used in each cycle of the loop, which (i) circumvents the possibly disruptive back action exerted by the measurement process on a quantum system, and (ii) permits the loop closure to be performed on laboratory apparatus cycling time scales (e.g., ∼10-3 s for liquid crystal laser modulators). The OCE learning process is based on the following realizations: (A) The molecule “knows” its own Hamiltonian, with no uncertainty. (B) When exposed to a laboratory control field, a molecule will solve its Schro¨dinger equation on ultrafast real molecular time scales, with absolute fidelity. (C) Laser pulse shapers with duty cycles of up to 104 distinct pulses per second are becoming available under full computer control. (D) Many physical objectives correspond to easily detectable outcomes, calling for little or no refined data analysis. (E) Fast algorithms exist to recognize patterns in the emerging control field f observable relationships, to automatically suggest new (better) control fields. The synthesis of steps (A),...,(E) produces an efficient closedloop learning procedure for teaching lasers to control quantum systems, and a schematic of this process is shown in Figure (2). The learning control procedure for manipulating quantum systems generally involves five basic elements: (1) an input trial control laser design, (2) the laboratory apparatus for generation of shaped laser pulses, (3) application of the laser control fields to the quantum system sample, (4) observation of the resultant control outcome, and (5) a learning algorithm that analyzes the measurement results from the prior experiments and suggests a new control field to be used in the next loop cycle. All of the current closed-loop learning control experiments1,30-32,35,88,89 were started by generating random initial control fields (i.e., side-stepping element (1) above). However, an OCT design may yield a good initial estimate for further laboratory OCE refinement as well as provide helpful guidance on the physical mechanism involved.73,87 An important enabling technology for laboratory quantum learning control is the ability to shape ultrafast laser pulses on the femtosecond scale.108 This technology is presently available and rapidly improving. Phase and amplitude modulation of the frequency components of the dispersed pulse is performed in the focal plane typically by an acousto-optic modulator (AOM)17 or by a liquid crystal modulator (LCM).20 The advantages of the AOM include high spectral resolution and fast response time, but it suffers from low light transmission (typically, about 5%). A LCM exhibits high light transmission (about 80%) and easy implementation (these devices are commercially available for the spectral range from 430 nm to 1.6 µm). However, a LCM has low spectral resolution (typically 128 discrete pixels) and slow transformation times, requiring at least a millisecond to change the pixels. Fast transformation times are important for closed-loop learning control, which may require exploring many thousands of distinct pulse shapes before finding an optimal result. However, current molecular implementations are not

significantly limited by the number of LCM pixels or the pixel transformation time. The present pulse-shaping technology in the visible and nearinfrared spectral ranges is suitable for exciting transitions between molecular electronic surfaces and for the manipulation of highly excited molecular vibrations. Further progress in the development of pulse shapers working in the mid- and farinfrared spectral ranges is necessary for control of molecules in their ground electronic state. The stability of pulse shapers appears to be adequate for the majority of chemical applications. The presence of modest noise in the laser electric field does not limit the quality of laboratory learning techniques employing evolutionary algorithms, and can even help the search for a better solution in a complex multidimensional parameter space.74,75,80 The opinion existed just a few years ago that the use of monochromatic lasers for control would be preferable in practice because laboratory learning control with tailored pulses would be too difficult to implement. However, recent advances in pulse-shaping technology have made this point of view obsolete, as demonstrated by closed-loop laboratory methods employing ultrashort shaped laser pulses that are rapidly becoming common experimental practice, with excellent reliability. There is no need to measure the laser field in the learning process, because any systematic characterization of the control “knobs” (e.g., pulse shaper parameters) is sufficient. This set of control knobs, determined by the experimental apparatus, defines the parameter space to be searched by the learning algorithm for an optimal laser shape. The closed-loop OCE procedure naturally incorporates any laboratory constraints on the control laser fields. Moreover, the algorithm will identify only those pathways to desired products that are adequately robust to inevitable random disturbances encountered in the laboratory.80 The learning algorithm should be sufficiently intelligent to ensure that the cyclic control process will converge on the objective. A variety of learning algorithms may be employed,75,76,79 but following the original proposal,2 the recent laboratory studies1,31-33,35,84,88,109-111 have focused on the use of evolutionary genetic algorithms26,112-114 as global search techniques. These techniques are quite effective even when noise is present,74,75,80 both in the measurements of the control outcome and in the tuning of the laser field. In contrast to excellent OCE performance, global search techniques are very difficult to use in numerical OCT simulations of learning control, because they demand enormous computational efforts.112 In such theoretical simulations, the global search for the optimum requires numerous iterations of the computationally intensive task of solving the time-dependent Schro¨dinger equation. However, in OCE learning control, the global search is remarkably efficient, because the evolving quantum system naturally “solves” its Schro¨dinger equation as accurately and as fast as possible. This OCE procedure eliminates the numerical burden of solving Schro¨dinger’s equation and allows for real-time adaptive control in the closedloop experiments. Although algorithms of the evolutionary type (e.g., genetic algorithms) are technically effective, they suffer from the conceptual drawback that little information about the control mechanisms is gained from the optimum search process. Recent progress has been made toward designing new algorithms for laboratory control that bring more insight into the physical processes operative in the controlled system.81 One approach uses high-dimensional model representation (HDMR), which provides a systematic means to experimentally determine the

Feature Article

J. Phys. Chem. A, Vol. 106, No. 27, 2002 6437

functional relationship between the applied control field and the resulting value of the control objective.115,116 In HDMR, the input (e.g., the control field) and the output (e.g., the control objective) are related through a hierarchy of control variable correlations. Nonlinear input f output maps based on HDMR have been recently employed to create an algorithm for laboratory learning control.81 This task is facilitated by expressing the control phases, amplitudes, or other laboratory parameters as a set of n input variables (x1,...,xn) ≡ x. A hierarchical map between a laboratory observable O and x may be written as

O(x) ) O0 +

Oij(xi,xj) + ... ∑i Oi(xi) + ∑ i